Selective Oxygenation of Cyclohexene by ... - ACS Publications

0 downloads 0 Views 667KB Size Report
Apr 19, 2017 - ABSTRACT: An iron complex with a tetraamido macrocyclic ligand, [(TAML)FeIII]. −. , was found to be an efficient and selective catalyst for ...
Article pubs.acs.org/IC

Selective Oxygenation of Cyclohexene by Dioxygen via an Iron(V)Oxo Complex-Autocatalyzed Reaction Muniyandi Sankaralingam,† Yong-Min Lee,† Wonwoo Nam,*,† and Shunichi Fukuzumi*,†,‡ †

Department of Chemistry and Nano Science, Ewha Womans University, Seoul 03760, Korea Faculty of Science and Engineering, Meijo University, SENTAN, Japan Science and Technology Agency (JST), Nagoya, Aichi 468-8502, Japan



S Supporting Information *

ABSTRACT: An iron complex with a tetraamido macrocyclic ligand, [(TAML)FeIII]−, was found to be an efficient and selective catalyst for allylic oxidation of cyclohexene by dioxygen (O2); cyclohex-2-enone was obtained as the major product along with cyclohexene oxide as the minor product. An iron(V)-oxo complex, [(TAML)FeV(O)]−, which was formed by activating O2 in the presence of cyclohexene, initiated the autoxidation of cyclohexene with O2 to produce cyclohexenyl hydroperoxide, which reacted with [(TAML)FeIII]− to produce [(TAML)FeV(O)]− by autocatalysis. Then, [(TAML)FeV(O)]− reacted rapidly with [(TAML)FeIII]− to produce a μ-oxo dimer, [(TAML)FeIV(O)FeIV(TAML)]2−, which was ultimately converted to [(TAML)FeV(O)]− when [(TAML)FeIII]− was not present in the reaction solution. An induction period was observed in the autocatalytic production of [(TAML)FeV(O)]−. The induction period was shortened with increasing catalytic amounts of [(TAML)FeV(O)]− and cyclohexenyl hydroperoxide, whereas the induction period was prolonged by adding catalytic amounts of a spin trapping reagent such as 5,5-dimethyl-1-pyrroline N-oxide (DMPO). The allylic oxidation of cycloalkenes was also found to depend on the allylic C−H bond dissociation energies, suggesting that the hydrogen atom abstraction from the allylic C−H bonds of cycloalkenes is the rate-determining radical chain initiation step. In this study, we have shown that an iron(III) complex with a tetraamido macrocyclic ligand is an efficient catalyst for the allylic oxidation of cyclohexene via an autocatalytic radical chain mechanism and that [(TAML)FeV(O)]− acts as a reactive intermediate for the selective oxygenation of cyclohexene with O2 to produce cyclohex2-enone predominantly.



INTRODUCTION Allylic oxidation reactions are fundamental and important for C−H bond functionalizations, accompanied by a significant increase in the synthetic and/or commercial value of the target molecules, because the resulting oxidation products are attractive synthetic intermediates for pharmaceuticals.1−3 Classically, allylic oxidations are performed with stoichiometric amounts of chromium reagents.3 Due to the high toxicity of the chromium reagents, however, the stoichiometric use of toxic oxidants should be avoided and replaced by environmentally friendly methods whenever possible.4−14 Among various oxidants, dioxygen (O2) is an ideal oxidant because it is ubiquitous and environmentally totally benign. Thus, a number of metal-catalyzed methods have been developed for allylic oxidations by O2.15−18 However, the catalytic oxidation of alkenes by O2 is normally not selective.18 For example, the catalytic oxidation of cyclohexene by O2 afforded mixtures of cyclohex-2-enol, cyclohex-2-enone, cyclohexene oxide, and other products, including cyclohexenyl hydroperoxide and 1,2-cyclohexanediol (Scheme 1).19 Thus, there has been no report on the selective, catalytic oxidation of cyclohexene by O2 © 2017 American Chemical Society

Scheme 1. Oxidation Products of Cyclohexene

under mild conditions. In addition, the mechanism of the catalytic oxidation of cyclohexene by O2 has yet to be clearly clarified. We report herein the catalytic and selective allylic oxidation of cyclohexene by O2 to produce cyclohex-2-enone as the major product together with cyclohexene oxide as the minor product in the presence of an iron complex bearing a tetraamido macrocyclic ligand, Na[(TAML)FeIII], in O2-saturated acetonitrile (MeCN) at 298 K (Scheme 2a for the structure of [(TAML)FeIII]− and Scheme 2b for the products formed in the autoxidation of cyclohexene by [(TAML)FeIII]−). The catalytic mechanism is proposed on the basis of a detailed kinetics study Received: January 30, 2017 Published: April 19, 2017 5096

DOI: 10.1021/acs.inorgchem.7b00220 Inorg. Chem. 2017, 56, 5096−5104

Article

Inorganic Chemistry Scheme 2. (a) Molecular Structure of Iron(III) Tetraamido Macrocyclic Complex ([(TAML)FeIII]−) and (b) Allylic Oxidation Products Formed in the Oxidation of Cyclohexene with [(TAML)FeIII]− in O2-Saturated Acetonitrile (MeCN) at 298 K

Table 1. Catalytic Oxidation of Cyclohexene by O2 in the Presence of [(TAML)FeIII]− in O2-Saturated Acetonitrilea

Reaction conditions: the catalytic activity of [(TAML)FeIII]− toward cyclohexene was examined by bubbling O2 into the reaction solution of deuterated acetonitrile (CD3CN) for 4 h at 298 K. Products were analyzed and quantified by taking 1H NMR spectra of the reaction solution. a

and the detection of intermediates in the catalytic oxidation reactions. This is the first time the catalytic and selective oxidation of cyclohexene by O2 has been achieved under ambient temperature and pressure with the elucidation of the autocatalytic reaction mechanisms.



RESULTS AND DISCUSSION Catalytic Oxidation of Cyclohexene by Dioxygen in the Presence of [(TAML)FeIII]− . The [(TAML)Fe III ]− complex with Na+ countercation20,21 exhibited no reactivity toward oxygen in acetonitrile, although the [(TAML)FeIII(H2O)]− complex with PPh4+ countercation was reported to be oxidized by oxygen to produce the μ-oxo dimer in dichloromethane.22 In this study, the oxidation reaction was carried out with Na[(TAML)FeIII] (2.0 mM) and cyclohexene (50 mM) in an O2-saturated MeCN solution at 298 K, yielding cyclohex-2-enone (90%) and cyclohexene oxide (10%) without the formation of cyclohex-2-enol, cyclohexenyl hydroperoxide, or cyclohexanediol (Figure 1 and Table 1). The solvent MeCN molecule may be coordinated to [(TAML)FeIII]−. The products were identified and quantified by 1H NMR as well as gas chromatography (GC) and gas chromatography linked to a mass spectrometer (GC-MS); the yields of products were within experimental errors (±10%) in the 1H NMR and GC

analyses (see Figure S1 in the Supporting Information for the analysis of products using 1H NMR). As shown in Table 1, an increase in the substrate concentration to 200 mM resulted in the formation of cyclohex-2-enone as the major product together with cyclohexene oxide and cyclohex-2-enol as the oxygenated products. Then, the decrease of the catalyst (20 μM) in an O2-saturated MeCN solution containing cyclohexene (200 mM) resulted in the formation of cyclohex-2-enone as the major product together with cyclohex-2-enol, cyclohexene oxide, and cyclohexenyl hydroperoxide as the oxygenated products. The formation of cyclohex-2-enone, cyclohex-2-enol, cyclohexenyl hydroperoxide, and cyclohexene oxide was confirmed by 1H NMR measurements (Figure S1 in the Supporting Information). The turnover number (TON) reached >104 by a decrease in the catalyst concentration, although the selectivity of the allylic oxidation products was somewhat diminished (Table 1). We have also examined the effect of concentrations of the catalyst and substrates (Table 1): the concentration of cyclohexenyl hydroperoxide decreased with increasing concentration of [(TAML)FeIII]−, because cyclohexenyl hydroperoxide reacts with [(TAML)FeIII]− to produce an iron(V)-oxo complex, [(TAML)FeV(O)]− (vide infra). In the absence of the [(TAML)FeIII]− catalyst, no formation of the oxygenated products was observed. In addition, iron salts (e.g., Fe(CF3SO3)2) and iron complexes, such as [Fe(TMC)]2+ (TMC = 1,4,8,11-tetramethyl-1,4,8,11-tetraazacyclotetradecane),23 [Fe(N4Py)]2+ (N4Py = N,N-bis(2-pyridylmethyl)-Nbis(pyridyl)methylamine),24 and [Fe(Bn-TPEN)]2+ (Bn-TPEN = N-benzyl-N,N′,N′-tris(2-pyridylmethyl)-1,2-diaminoethane),25 did not yield the oxygenated products under the catalytic conditions described above. These results demonstrate that [(TAML)FeIII]− is a highly efficient catalyst for the oxidation of cyclohexene by O2. We have also performed the catalytic reaction in the presence of isotopically labeled H218O.26 In the reaction mixture, nearly 50% incorporation of

Figure 1. Reaction time profiles for the formation of cyclohex-2-enone (green) and cyclohexene oxide (red) and for the consumption of cyclohexene (black) in the oxidation of cyclohexene (50 mM) in the presence of [(TAML)FeIII]− (2.0 mM) in O2-saturated MeCN at 298 K. 5097

DOI: 10.1021/acs.inorgchem.7b00220 Inorg. Chem. 2017, 56, 5096−5104

Article

Inorganic Chemistry 18

Supporting Information), whose mass and isotope distribution pattern corresponds to [(TAML)FeV(O)]− (calculated m/z 442.1). When the reaction was carried out with isotopically labeled H218O, a mass peak appeared corresponding to [(TAML)FeV(18O)]− (m/z 444.1) (Figure S4), indicating that the intermediate contains only one oxygen atom. Thus, the stability of [(TAML)FeV(O)]− at 298 K increased under our reaction conditions, such as using O2 as an oxidant. It should be also noted that the [(TAML)FeV(O)]− complex generated by other oxidants, such as cumene hydroperoxide and cyclohexenyl hydroperoxide, at 298 K (see Figures S5 and S6 in the Supporting Information) is much more stable than that produced by m-CPBA,21,27 probably indicating that the presence of H+ derived from m-CPBA may decrease the stability of [(TAML)FeV(O)]−. Addition of a catalytic amount of [(TAML)FeV(O)]− to an MeCN solution of [(TAML)FeIII]− containing cyclohexene resulted in a significant decrease in the induction period, which became shorter with an increase in the catalytic amount of [(TAML)FeV(O)]− (Figure 3a). Because [(TAML)FeV(O)]− is the product of the oxidation of [(TAML)FeIII]− with O2 in the presence of cyclohexene, the decrease in the induction period in Figure 3a indicates an autocatalytic behavior; an autocatalytic reaction is defined as one in which the product acts as the catalyst for its own formation.30 Because the rate of reaction is proportional to the concentration of product, the time course of the reaction exhibits an induction period with a sigmoidal curvature, as shown in Figure 3a.30 Autocatalysis has been extensively studied because it is central to the propagation of living systems.31−38 Addition of a catalytic amount of the μ-oxo dimer ([(TAML)FeIV(O)FeIV(TAML)]2−) to an MeCN solution of [(TAML)FeIII]− containing cyclohexene decreased the induction period in a similar way (Figure 3b). The induction period was also reduced by addition of a catalytic amount of cyclohexenyl hydroperoxide (Figure 3c), which was independently prepared by photoinduced autoxidation of cyclohexene by O2 with xyloquinone,39 and the concentration was determined on the basis of iodometric titration (Figures S1e and S7 in the Supporting Information).40 Addition of a catalytic amount of cumene hydroperoxide also resulted in the reduction of the induction period with an increase in the concentration of cumene hydroperoxide (Figure S8 in the Supporting Information), because cumene hydroperoxide reacted with [(TAML)FeIII]− to produce [(TAML)FeV(O)]− (Figure S5 in the Supporting Information). The induction period was also reduced by increasing the concentration of cyclohexene (Figure 4a). In contrast, the induction period increased with increasing concentration of [(TAML)FeIII]− (Figure 4b). Since the [(TAML)FeV(O)]− complex rapidly reacts with [(TAML)FeIII]− to yield [(TAML)FeIV(O)FeIV(TAML)]2−, which is in equilibrium with [(TAML)FeV(O)]− and the equilibrium shifted to the right-hand side for the formation of the μ-oxo dimer,41,42 the concentration of [(TAML)FeV(O)]−, which acted as the autocatalyst, decreased with increasing concentration of [(TAML)FeIII]− when the induction period increased (Figure 4b). These autocatalytic behaviors suggest that a radical chain autoxidation is responsible for the formation of [(TAML)FeIV(O)FeIV(TAML)]2− and [(TAML)FeV(O)]− in the oxidation of [(TAML)FeIII]− by O2 in the presence of cyclohexene, as shown in Scheme 3 for the autocatalytic radical chain reaction induced by the [(TAML)FeV(O)]− complex (see below).

O-labeled cyclohexene oxide was observed (Figure S2 in the Supporting Information). Autocatalytic Formation of [(TAML)FeV(O)]− in the Reaction of [(TAML)FeIII] − with Cyclohexene and Dioxygen. Addition of [(TAML)FeIII]− (0.10 mM) to an O2-saturated MeCN solution containing cyclohexene (2.5 mM) resulted in UV−vis spectral changes of the iron complex (Figure 2a), where the absorption band at 400 nm due to

Figure 2. UV−visible absorption spectral changes observed in the oxygenation of [(TAML)FeIII]− (0.10 mM) with O2 in the presence of cyclohexene (2.5 mM) in O2-saturated MeCN at 298 K. The two-step changes are separated by the time regions of (a) 0−720 s and (b) 720−800 s.

[(TAML)FeIII]− decreased with isosbestic points at 372 and 415 nm, accompanied by the formation of a μ-oxo dimer, [(TAML)FeIV(O)FeIV(TAML)]2−,27−29 with an induction period (Figure 2a, inset). The latter species was then converted rapidly to an iron(V)-oxo complex, [(TAML)FeV(O)]−, with isosbestic points at 572 and 692 nm in the time region of 720− 780 s (Figure 2b). The yield of the iron(V)-oxo complex was determined to be >95% on the basis of the extinction coefficient at 630 nm.21,27−29 The formation of the iron(V)oxo complex has also been confirmed by electron paramagnetic resonance (EPR) spectroscopy and cold spray ionization timeof-flight mass (CSI-MS) spectrometry. The major anisotropic EPR signals with gxx = 1.98, gyy = 1.97, and gzz = 1.76 in the EPR spectrum recorded in acetone at 77 K belong to S = 1/2 [(TAML)FeV(O)]− species (Figure S3 in the Supporting Information). A simulated EPR spectrum using the anisotropic g-tensors of gxx = 1.979, gyy = 1.973, and gzz = 1.763 was also represented (Figure S3a, pink dotted line). The spin amount of the iron(V)-oxo complex (78(4)%) was determined by a comparison of the doubly integrated value of the EPR signal with that of 2,2-diphenyl-1-picrylhydrazyl radical (DPPH•) species as a reference (Figure S3). The signals with g values of 2.06 and 2.03 originated from the minor species.27 The CSI-MS spectrum of [(TAML)FeV(O)]− exhibits a prominent ion peak at a mass to charge ratio of m/z 442.1 (Figure S4 in the 5098

DOI: 10.1021/acs.inorgchem.7b00220 Inorg. Chem. 2017, 56, 5096−5104

Article

Inorganic Chemistry

Figure 4. (a) Time profiles monitored for the change of absorbance at 630 nm due to [(TAML)FeV(O)]− in the autocatalytic oxygenation of [(TAML)FeIII]− (0.10 mM) with O2 in the presence of cyclohexene (2.5−10.0 mM). (b) Time profiles monitored for the change of absorbance at 630 nm due to [(TAML)FeV(O)]− in the reaction of cyclohexene (5.0 mM) with [(TAML)FeIII]− (0.10−0.25 mM) in aerated MeCN at 298 K.

Scheme 3. Autocatalytic Radical Chain Mechanism for Catalytic Oxidation of Cyclohexene by O2 in the Presence of [(TAML)FeIII]− Figure 3. Time profiles monitored for the change of absorbance at 630 nm due to [(TAML)FeV(O)]− in the autocatalytic oxygenation of [(TAML)FeIII]− (0.10 mM) with O2 in the presence of cyclohexene (2.5 mM) with catalytic amounts of (a) [(TAML)FeV(O)]− (0−10%), (b) [(TAML)FeIV(O)FeIV(TAML)]2− (0−15%), and (c) cyclohexenyl hydroperoxide (0−1.5%) in aerated MeCN at 298 K.

It has been known that the initiation of autoxidation of cyclohexene by O2 occurs without any catalyst.43 The radical chain reaction is then started by hydrogen atom abstraction of the allylic C−H bonds of cyclohexene by [(TAML)FeV(O)]− to produce [(TAML)FeIV(OH)]− and cyclohexenyl radical (Scheme 3, reaction a), which readily reacts with O2 to afford cyclohexenylperoxyl radical (Scheme 3, reaction b). The produced [(TAML)FeIV(OH)]− may be dimerized to produce the μ-oxo dimer. The cyclohexenylperoxyl radical is the chain carrier, which abstracts a hydrogen atom from cyclohexene to produce cyclohexenyl hydroperoxide and the regeneration of cyclohexenyl radical (Scheme 3, reaction c), constituting the radical chain reactions.44 Cyclohexenyl hydroperoxide oxygenates [(TAML)Fe III ] − to produce cyclohex-2-enol and [(TAML)FeV(O)]− (Scheme 3, reaction d). It was confirmed that addition of cyclohexenyl hydroperoxide to an MeCN solution of [(TAML)FeIII]− resulted in the generation of [(TAML)FeV(O)]− (Figure S6 in the Supporting Information). In the presence of [(TAML)FeIII]−, [(TAML)FeV(O)]− is known to react rapidly with [(TAML)FeIII]− to produce the μ-

oxo dimer [(TAML)FeIV(O)FeIV(TAML)]2− (Scheme 3, reaction e).21,27−29,41,42 When [(TAML)FeIII]− was completely consumed, [(TAML)FeIV(O)FeIV(TAML)]2− was rapidly converted to [(TAML)FeV(O)]− (Scheme 3, reaction f), as shown in Figure 2b. The overall chain reaction is given in eq 1, where cyclohexene reacts with O2 and [(TAML)FeIII]− to yield [(TAML)FeV(O)]−. Then, cyclohexene is directly oxygenated by [(TAML)FeV(O)]− to produce cyclohex-2-enol (Scheme 4, reaction a), which is further oxidized by [(TAML)FeV(O)]− to 5099

DOI: 10.1021/acs.inorgchem.7b00220 Inorg. Chem. 2017, 56, 5096−5104

Article

Inorganic Chemistry

Scheme 4. Stepwise Oxidation of Cyclohexene with [(TAML)FeIII]− (Major Pathway)

Figure 5. (a) UV−visible spectral changes observed in the reaction of [(TAML)FeV(O)]− (0.20 mM) with cyclohex-2-enol (5.0 mM) in MeCN at 233 K. [(TAML)FeV(O)]− was generated by the O2activation reaction of [(TAML)FeIII]− (0.20 mM) with cyclohexene (5.0 mM) in O2-saturated MeCN. (b) UV−visible spectral changes observed in the reaction of [(TAML)FeV(O)]− (0.20 mM) with cyclohexene (5.0 mM) in O2-saturated MeCN at 233 K. Insets show the time courses monitored at 840 nm.

yield cyclohex-2-enone (Scheme 4, reaction b), accompanied by the regeneration of [(TAML)FeIII]−, which is in equilibrium with [(TAML)FeIV(O)FeIV(TAML)]2− in the presence of [(TAML)FeV(O)]− (Scheme 4). The rapid oxidation of cyclohex-2-enol by [(TAML)FeV(O)]− was independently confirmed (Figure 5a; see also Figure S9 in the Supporting Information), in which the rate of the reaction of [(TAML)FeV(O)]− with cyclohex-2-enol was about 10 times faster than that of [(TAML)FeV(O)]− with cyclohexene at 233 K (Figure 5b). As a minor reaction pathway, cyclohexene was epoxidized by [(TAML)FeV(O)]− to give cyclohexene epoxide (Scheme 5). Once the autocatalytic radical chain reaction starts, the initiation step by [(TAML)Fe V (O)]− produced in an autocatalytic pathway (Scheme 3) becomes dominant in comparison with the O2 activation pathway. Previously we have reported the formation of [FeIV(O)(TMC)]2+ by reacting [FeII(TMC)]2+ with dioxygen in the presence of cycloalkenes.23c In the case of dioxygen activation by [FeII(TMC)]2+, however, no autocatalytic pathway was involved, in contrast to the present study of [(TAML)FeIII]−. The UV−vis spectral changes were also monitored under the same conditions as in Figure 1 to clarify the rate-determining step in Scheme 3. The results are shown in Figure S10 in the Supporting Information, where spectral changes were observed similar to those seen in Figure 2. When the cyclohexene concentration was 50 mM, which was much larger than that in Figure 2 (2.5 mM), no induction period was observed (Figure S10a). The nearly complete formation of [(TAML)FeV(O)]− via the formation of [(TAML)FeIV(O)FeIV(TAML)]2− from [(TAML)FeV(O)]− and [(TAML)FeIII]− indicates that the oxygenation of cyclohexene with [(TAML)FeV(O)]− is the rate-determining step (Scheme 3, reaction a). After the completion of the oxidation of cyclohexene, [(TAML)FeIII]− was regenerated (Figure S10b). The regenerated [(TAML)FeIII]− was confirmed by EPR and CSI-MS (Figure S11 in the Supporting Information). Inhibition of the Radical Chain Reaction. The chain carrier radical (cylcohexenylperoxyl radical) was trapped by

Scheme 5. Epoxidation of Cyclohexene with [(TAML)FeIII]− (Minor Pathway)

5,5′-dimethyl-1-pyrroline N-oxide (DMPO), which is a typical spin-trapping reagent.45−47 The addition of only 0.001% DMPO to an aerated MeCN solution of [(TAML)FeIII]− and cyclohexene resulted in an increase in the induction period, and the addition of 2% DMPO inhibited the reaction completely, as shown in Figure 6. Such a strong inhibition of the reaction by DMPO suggests a long radical chain length. Since the rate constants of the spin-trapping reactions of peroxyl radicals with DMPO (103−104 M−1 s−1)48 are much larger than those of allylic hydrogen atom abstraction from alkenes by peroxyl radicals (e.g., 2.7 M−1 s−1 for the reaction of tert-butylperoxyl radical with tetralin),49,50 the rate of consumption of DMPO in the induction period ([DMPO]/(t4 − t2)) corresponds to the initiation rate. The propagation rate is obtained as the fast rate after the induction period (i.e., [[(TAML)FeIII]−]/(t3 − t1)) (see Figure 7). Then, the chain length of the propagation step in Scheme 3 is evaluated as the ratio of the propagation rate to the initiation rate, {[[(TAML)FeIII]−]/(t3 − t1)}/{[DMPO]/ 5100

DOI: 10.1021/acs.inorgchem.7b00220 Inorg. Chem. 2017, 56, 5096−5104

Article

Inorganic Chemistry

Figure 6. Time profiles monitored for the change of absorbance at 630 nm due to [(TAML)FeV(O)]− in the oxidation of [(TAML)FeIII]− (0.10 mM) with O2 in the presence of cyclohexene (2.5 mM) with [(TAML)FeIII]− (0.10 mM) and a catalytic amount of DMPO (0− 2.0%) with O2 in aerated MeCN at 298 K.

Figure 8. (a) UV−vis spectral changes observed in the reversible interconversion between [(TAML)FeV(O)]− and [(TAML)FeIV(O)FeIV(TAML)]2− by changing the temperature between 253 and 233 K. Initially, [(TAML)FeV(O)]− was generated by an O2 activation reaction of [(TAML)FeIII]− (0.20 mM) with cyclohexene (5.0 mM) in O2-saturated MeCN at 298 K. (b) Time profiles monitored at 840 nm due to [(TAML)FeIV(O)FeIV(TAML)]2− by changing the reaction temperature between 253 and 233 K. Figure 7. Chain length evaluated as the ratio of the propagation rate to the initiation rate. The initiation rate is equal to [DMPO]/(t4 − t2), whereas the propagation rate is equal to [[(TAML)FeIII]−]/(t3 − t1): t1 = 580 s; t2 = 630 s; t3 = 830 s; t4 = 2240 s. The chain length was calculated from {[[(TAML)FeIII]−]/(t3 − t1)}/{[DMPO]/(t4 − t2)} to be 650.

(Scheme 3). When the temperature was increased to 253 K, however, the absorbance at 840 nm due to [(TAML)FeIV(O)FeIV(TAML)]2− decreased, accompanied by an increase in absorbance at 630 nm due to [(TAML)FeV(O)]−. This result indicates that, at a higher temperature, the autocatalytic formation of [(TAML)FeV(O)]− (Scheme 3) is faster than the oxidation of cyclohexene by [(TAML)FeV(O)]− (Schemes 4 and 5). Such interconversion between [(TAML)FeV(O)]− and [(TAML)FeIV(O)FeIV(TAML)]2− can be repeated by changing the temperature between 233 and 253 K (Figure 8). Catalytic Oxygenation of Other Cycloalkenes by O2. Other cycloalkenes, such as cycloheptene, were also oxygenated by O2 in the presence of a catalytic amount of [(TAML)FeIII]− to produce the corresponding allylic oxidation products with epoxides (Table 2). In the case of cyclooctene and deuterated

(t4 − t2)}, which is ca. 650 (Figure 7). Thus, the chain length is long enough to obtain large TONs in Table 1. It should be noted that the autocatalytic radical chain reaction is for the formation of [(TAML)FeV(O)]−, which acts as an efficient catalyst for the selective oxidation of cyclohexene to cyclohex-2enone as the major oxidized product, in contrast with the autoxidation pathway to produce cyclohexenyl hydroperoxide. Interconversion between [(TAML)Fe V (O)] − and [(TAML)FeIV(O)FeIV(TAML)]2− Depending on Temperature. The balance for the autocatalytic formation of [(TAML)FeV(O)]− in Scheme 3 and the oxidation of cyclohexene with [(TAML)FeV(O)]− (eq 1) is changed depending on the temperature. When the reaction temperature was lowered to 233 K after the formation of [(TAML)FeV(O)]− at 298 K, the absorbance at 630 nm due to [(TAML)FeV(O)]− decreased, accompanied by the appearance of an absorption band at 840 nm due to the formation of the μoxo dimer [(TAML)FeIV(O)FeIV(TAML)]2− in the reaction of [( TAM L)Fe V (O)] − a nd [ (T A M L )Fe I I I ] − (Figure 8);21,27−29,41,42 cyclohexene is oxidized by [(TAML)FeV(O)]− to produce cyclohex-2-enol and [(TAML)FeIII]−, which reacts with [(TAML)FeV(O)]− rapidly to yield the μ-oxo dimer species [(TAML)FeIV(O)FeIV(TAML)]2− (Schemes 4 and 5). The reaction of [(TAML)FeV(O)]− and [(TAML)FeIII]− is faster than the autocatalytic formation of [(TAML)FeV(O)]−

Table 2. Products Formed in the Catalytic Oxidation of Cycloalkenes by O2 in the Presence of [(TAML)FeIII]− in O2-Saturated Acetonitrile at 298 Ka product (%) substrate

alcohol

ketone

epoxide

cyclohexene cycloheptene cyclooctene

0 10(1) 0

89(2) 22(2) 0

11(1) 22(1) 0

a

Reaction conditions: the oxygenation of cycloalkenes was performed by bubbling O2 into a CD3CN solution containing [(TAML)FeIII]− (2.0 mM) and cycloalkenes (50 mM) at 298 K for 4 h. Products were analyzed and quantified by taking 1H NMR spectra and comparing them with those of authentic samples. 5101

DOI: 10.1021/acs.inorgchem.7b00220 Inorg. Chem. 2017, 56, 5096−5104

Article

Inorganic Chemistry

collected on a JMS-T100CS (JEOL) mass spectrometer equipped with a CSI source. Typical measurement conditions are as follows: needle voltage, 2.2 kV; orifice 1 current, 50−500 nA; orifice 1 voltage, 0−20 V; ring lens voltage, 10 V; ion source temperature, 5 °C; spray temperature, −40 °C. The product analyses were also performed with an Agilent 6890N gas chromatograph (GC) and a FOCUS DSQ (dual-stage quadrupole) mass spectrometer (Thermo Finnigan, Austin, TX, USA) interfaced with a Finnigan FOCUS gas chromatograph (GC-MS). Formation of Fe(V)-Oxo Intermediate. Reactions were run in a 1 cm UV cuvette by monitoring UV−vis spectral changes of reaction solutions. Addition of a catalytic amount of [(TAML)FeIII]− (0.10 mM) to an O2-saturated MeCN solution containing cyclohexene (2.5 mM) resulted in the formation of [(TAML)FeV(O)]− at 298 K in a good yield, which was determined using the extinction coefficient at 630 nm due to [(TAML)FeV(O)]−. An authentic [(TAML)FeV(O)]− complex was synthesized by reacting [(TAML)FeIII]− with mchloroperbenzoic acid (m-CPBA) according to the published procedure.27 Catalytic Oxygenation of Cycloalkenes by O2 in the Presence of [(TAML)FeIII]−. The catalytic activity of [(TAML)FeIII]− toward cycloalkenes was examined by bubbling dioxygen into the reaction solution of acetonitrile (MeCN or CD3CN). Oxygenated products were identified by 1H NMR and GC and then compared with those of authentic samples to determine product yields. Synthesis and Characterization of Cyclohexenyl Hydroperoxide. Cyclohexenyl hydroperoxide was generated using cyclohexene (450 μL, 9.8 M) and xyloquinone (2.0 mM) after 15 h of irradiation using a xenon lamp (300 W) at 298 K according to the literature method for the photocatalytic production of hydroperoxides.39 It was confirmed by 1H NMR spectroscopy; the hydroperoxide peaks match with the previously reported literature.54 Iodometric Titration of Cyclohexenyl Hydroperoxide. The amount of cyclohexenyl hydroperoxide was determined by titration with iodide ion;40 10 μL of the reaction solution containing cyclohexenyl hydroperoxide generated by cyclohexene (450 μL, 9.8 M) and xyloquinone (2.0 mM) with irradiation was taken and diluted up to 2.0 mL with MeCN. The reaction solution was then treated with an excess of sodium iodide under an Ar atmosphere. The amount of I3− formed was quantified by absorbance at 361 nm due to I3− (λmax 361 nm, ε = 2.5 × 104 M−1 cm−1).40

cyclohexene (cyclohexene-d10), however, the oxidation by O2 with [(TAML)FeIII]− did not occur under the identical reaction conditions (Table 2). No [(TAML)FeV(O)]− was produced in the reaction of [(TAML)FeIII]− (0.10 mM) with O2 in the presence of cyclooctene (50 mM) in O2-saturated MeCN at 298 K for 4 h due to the much slower autoxidation to start the autocatalytic reaction (Figure S12 in the Supporting Information). The reactivity of cycloalkenes was correlated with the allylic C−H bond dissociation energies (BDEs) of cycloalkenes, since the yields of oxygenated products decrease with an increase in the C−H BDEs, such as cyclohexene (100%, BDE = 81.0 kcal mol−1), cycloheptene (54%, BDE = 83 kcal mol−1), and cyclooctene (0%, BDE = 85.0 kcal mol−1) (Table 2).51 Thus, the hydrogen atom abstraction from cycloalkenes by [(TAML)FeV(O)]− is the initiation step in the autoxidation radical chain reactions, as proposed in Scheme 3.



CONCLUSION The selective catalytic oxidation of cyclohexene by O2 has been achieved by using an iron complex bearing a tetraamido macrocyclic ligand, [(TAML)FeIII]−, which is oxidized by O2 in the presence of cyclohexene to yield [(TAML)FeV(O)]− and the corresponding allylic oxidation products via an autocatalytic radical chain mechanism.52 The nonselectivity in the oxidized products in Scheme 1 results mainly from free radical decomposition reactions of the autoxidation product (i.e., cyclohexenyl hydroperoxide). Cyclohexenyl hydroperoxide rapidly oxygenates [(TAML)FeIII]− to produce [(TAML)FeV(O)]−, which can selectively oxygenate cyclohexene to yield mainly cyclohex-2-enone, which would otherwise be difficult to obtain selectively. Thus, the present study provides a unique autocatalytic reaction mechanism for the selective catalytic oxygenation of substrates by O2 without the involvement of free radical decomposition of alkyl hydroperoxides.





EXPERIMENTAL SECTION

Materials. Commercially available chemicals and solvents were used without further purification unless otherwise indicated. The complex Na(H2O)x[(TAML)FeIII] was purchased from GreenOx Catalyst, Inc. (Pittsburgh, PA, USA), and recrystallized from an isopropyl alcohol/H2O mixture before use.20 Acetonitrile, cyclohexene, cyclohex-2-enone, cyclohex-2-enol, cyclohexene oxide, cycloheptene, cyclooctene, and 5,5′-dimethyl-1-pyrroline N-oxide were purchased from Aldrich Chemical Co. 2,5-Dimethyl-4-benzoquinone was purchased from Alfa-Aesar Chemical Co. Olefins were refluxed and distilled under Ar and filtered through an active alumina column prior to use.53 H218O (95% 18O enriched) was purchased from ICON Services Inc. (Summit, NJ, USA). Instrumentation. UV−vis spectra were recorded on a HewlettPackard 8453 diode array spectrophotometer equipped with a UNISOKU Scientific Instruments USP-203A Cryostat for lowtemperature experiments. X-band electron paramagnetic resonance (EPR) spectra were recorded at 5 K using an X-band Bruker EMXplus spectrometer equipped with a dual mode cavity (ER 4116DM). Low temperature was achieved and controlled with an Oxford Instruments ESR900 liquid He quartz cryostat with an Oxford Instruments ITC503 temperature and gas flow controller. The experimental parameters for EPR measurement were as follows: microwave frequency, 9.647 GHz; microwave power, 1.0 mW; modulation amplitude, 10 G; gain, 1 × 104; modulation frequency, 100 kHz; time constant, 40.96 ms; conversion time, 81.00 ms. Nuclear magnetic resonance (NMR) spectra were measured with a Bruker model digital AVANCE III 400 FT-NMR spectrometer. The cold spray ionization time-of-flight mass (CSI-MS) spectral data were

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.7b00220. Figures S1−S12 as described in the text (PDF)



AUTHOR INFORMATION

Corresponding Authors

*E-mail for W.N.: [email protected]. *E-mail for S.F.: [email protected]. ORCID

Yong-Min Lee: 0000-0002-5553-1453 Wonwoo Nam: 0000-0001-8592-4867 Shunichi Fukuzumi: 0000-0002-3559-4107 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported by the NRF of Korea through CRI (NRF-2012R1A3A2048842 to W.N.) and GRL (NRF-201000353 to W.N.), a Grant-in-Aid (no. 16H02268 to S.F.) from the Ministry of Education, Culture, Sports, Science and 5102

DOI: 10.1021/acs.inorgchem.7b00220 Inorg. Chem. 2017, 56, 5096−5104

Article

Inorganic Chemistry

Manganese(II) Complexes of Tetradentate 4N Ligands with Diazapane Backbones for the Catalytic Olefin Epoxidation: Effect of Nucleophilicity of Peroxo Complexes on Reactivity. RSC Adv. 2014, 4, 12000. (15) (a) Kojima, T.; Nakayama, K.; Sakaguchi, M.; Ogura, T.; Ohkubo, K.; Fukuzumi, S. Photochemical Activation of Ruthenium(II)-Pyridylamine Complexes Having a Pyridine-N-Oxide Pendant toward Oxygenation of Organic Substrates. J. Am. Chem. Soc. 2011, 133, 17901. (b) Hirai, Y.; Kojima, T.; Mizutani, Y.; Shiota, Y.; Yoshizawa, K.; Fukuzumi, S. Ruthenium-Catalyzed Selective and Efficient Oxygenation of Hydrocarbons with Water as an Oxygen Source. Angew. Chem., Int. Ed. 2008, 47, 5772. (16) (a) Kojima, T.; Matsuda, Y. Catalytic Hydrocarbon Oxygenation by a Dinuclear Ruthenium(II) Complex with Molecular Oxygen. Chem. Lett. 1999, 28, 81. (b) Ohzu, S.; Ishizuka, T.; Hirai, Y.; Jiang, H.; Sakaguchi, M.; Ogura, T.; Fukuzumi, S.; Kojima, T. Mechanistic Insight into Catalytic Oxidations of Organic Compounds by Ruthenium(IV)-Oxo Complexes with Pyridylamine Ligands. Chem. Sci. 2012, 3, 3421. (17) (a) Qi, Y.; Luan, Y.; Yu, J.; Peng, X.; Wang, G. Nanoscaled Copper Metal-Organic Framework (MOF) Based on Carboxylate Ligands as an Efficient Heterogeneous Catalyst for Aerobic Epoxidation of Olefins and Oxidation of Benzylic and Allylic Alcohols. Chem. - Eur. J. 2015, 21, 1589. (b) Wang, J.; Yang, M.; Dong, W.; Jin, Z.; Tang, J.; Fan, S.; Lu, Y.; Wang, G. Co(II) Complexes Loaded into Metal-Organic Frameworks as Efficient Heterogeneous Catalysts for Aerobic Epoxidation of Olefins. Catal. Sci. Technol. 2016, 6, 161. (18) Cao, Y.; Yu, H.; Peng, F.; Wang, H. Selective Allylic Oxidation of Cyclohexene Catalyzed by Nitrogen-Doped Carbon Nanotubes. ACS Catal. 2014, 4, 1617. (19) (a) Zou, G.; Jing, D.; Zhong, W.; Zhao, F.; Mao, L.; Xu, Q.; Xiao, J.; Yin, D. A Novel Route for Preparation of Mn-Containing Hollow Framework TS-1, and Its Selective Allylic Oxidation of Cyclohexene. RSC Adv. 2016, 6, 3729. (b) Boudjema, S.; Vispe, E.; Choukchou-Braham, A.; Mayoral, J. A.; Bachir, R.; Fraile, J. M. Preparation and Characterization of Activated Montmorillonite Clay Supported 11-Molybdo-Vanado-Phosphoric Acid for Cyclohexene Oxidation. RSC Adv. 2015, 5, 6853. (20) (a) Horwitz, C. P.; Ghosh, A. Carnegie Mellon University, Pittsburgh, PA, USA, 2006; found under http://igs.chem.cmu.edu/. (b) Collins, T. J.; Powell, R. D.; Slebodnick, C.; Uffelman, E. S. Stable Highly Oxidizing Cobalt Complexes of Macrocyclic Ligands. J. Am. Chem. Soc. 1991, 113, 8419. (21) Kwon, E.; Cho, K.-B.; Hong, S.; Nam, W. Mechanistic Insight into the Hydroxylation of Alkanes by a Nonheme Iron(V)-Oxo Complex. Chem. Commun. 2014, 50, 5572. (22) Ghosh, A.; Tiago de Oliveira, F.; Yano, T.; Nishioka, T.; Beach, E. S.; Kinoshita, I.; Münck, E.; Ryabov, A. D.; Horwitz, C. P.; Collins, T. J. Catalytically Active μ-Oxodiiron(IV) Oxidants from Iron(III) and Dioxygen. J. Am. Chem. Soc. 2005, 127, 2505. (23) (a) Rohde, J.-U.; In, J.-H.; Lim, M. H.; Brennessel, W. W.; Bukowski, M. R.; Stubna, A.; Münck, E.; Nam, W.; Que, L., Jr. Crystallographic and Spectroscopic Characterization of a Nonheme Fe(IV)O Complex. Science 2003, 299, 1037. (b) Hong, S.; Lee, Y.M.; Shin, W.; Fukuzumi, S.; Nam, W. Dioxygen Activation by Mononuclear Nonheme Iron(II) Complexes Generates Iron-Oxygen Intermediates in the Presence of an NADH Analogue and Proton. J. Am. Chem. Soc. 2009, 131, 13910. (c) Lee, Y.-M.; Hong, S.; Morimoto, Y.; Shin, W.; Fukuzumi, S.; Nam, W. Dioxygen Activation by a NonHeme Iron(II) Complex: Formation of an Iron(IV)-Oxo Complex via C-H Activation by a Putative Iron(III)-Superoxo Species. J. Am. Chem. Soc. 2010, 132, 10668. (24) (a) Lubben, M.; Meetsma, A.; Wilkinson, E. C.; Ferringa, B.; Que, L., Jr. Nonheme Iron Centers in Oxygen Activation: Characterization of an Iron(III) Hydroperoxide Intermediate. Angew. Chem., Int. Ed. Engl. 1995, 34, 1512. (b) Lee, Y.-M.; Kotani, H.; Suenobu, T.; Nam, W.; Fukuzumi, S. Fundamental Electron-Transfer Properties of Non-Heme Oxoiron(IV) Complexes. J. Am. Chem. Soc. 2008, 130, 434.

Technology (MEXT), and a SENTAN project from the Japan Science and Technology Agency (JST) to S.F.



REFERENCES

(1) (a) Weidmann, V.; Maison, W. Allylic Oxidations of Olefins to Enones. Synthesis 2013, 45, 2201. (b) Roduner, E.; Kaim, W.; Sarkar, B.; Urlacher, V. B.; Pleiss, J.; Gläser, R.; Einicke, W.-D.; Sprenger, G. A.; Beifuß, U.; Klemm, E.; Liebner, C.; Hieronymus, H.; Hsu, S.-F.; Plietker, B.; Laschat, S. Selective Catalytic Oxidation of C-H Bonds with Molecular Oxygen. ChemCatChem 2013, 5, 82. (2) Horn, E. J.; Rosen, B. R.; Chen, Y.; Tang, J.; Chen, K.; Eastgate, M. D.; Baran, P. S. Scalable and Sustainable Electrochemical Allylic CH Oxidation. Nature 2016, 533, 77. (3) (a) Nakamura, A.; Nakada, M. Allylic Oxidations in Natural Product Synthesis. Synthesis 2013, 45, 1421. (b) Garcia-Cabeza, A. L.; Moreno-Dorado, F. J.; Ortega, M. J.; Guerra, F. M. Copper-Catalyzed Oxidation of Alkenes and Heterocycles. Synthesis 2016, 48, 2323. (4) (a) Whitmore, F. C.; Pedlow, G. W., Jr. Δ2-Cyclohexenone and Related Substances. J. Am. Chem. Soc. 1941, 63, 758. (b) Cainelli, G.; Cardille, G. Chromium Oxidations in Organic Chemistry; Springer Verlag: Berlin, 1984. (5) (a) Jiang, D.; Mallat, T.; Meier, D. M.; Urakawa, A.; Baiker, A. Copper Metal-Organic Framework: Structure and Activity in the Allylic Oxidation of Cyclohexene with Molecular Oxygen. J. Catal. 2010, 270, 26. (b) Dali, A.; Rekkab-Hammoumraoui, I.; ChoukchouBraham, A.; Bachir, R. Allylic Oxidation of Cyclohexene over Ruthenium-Doped Titanium-Pillared Clay. RSC Adv. 2015, 5, 29167. (6) Leus, K.; Vanhaelewyn, G.; Bogaerts, T.; Liu, Y.-Y.; Esquivel, D.; Callens, F.; Marin, G. B.; van Speybroeck, V.; Vrielinck, H.; van der Voort, P. Ti-Functionalized NH2-MIL-47: An Effective and Stable Epoxidation Catalyst. Catal. Today 2013, 208, 97. (7) Tuci, G.; Giambastiani, G.; Kwon, S.; Stair, P. C.; Snurr, R. Q.; Rossin, A. Chiral Co(II) Metal-Organic Framework in the Heterogeneous Catalytic Oxidation of Alkenes under Aerobic and Anaerobic Conditions. ACS Catal. 2014, 4, 1032. (8) Ruano, D.; Díaz-García, M.; Alfayate, A.; Sánchez-Sánchez, M. Nanocyrstalline M-MOF-74 as Heterogeneous Catalysts in the Oxidation of Cyclohexene: Correlation of the Activity and Redox Potential. ChemCatChem 2015, 7, 674. (9) Xu, L.; Huang, D.-D.; Li, C.-G.; Ji, X.; Jin, S.; Feng, Z.; Xia, F.; Li, X.; Fan, F.; Li, C.; Wu, P. Construction of Unique Six-Coordinated Titanium Species with an Organic Amine Ligand in Titanosilicate and their Unprecedented High Efficiency for Alkene Epoxidation. Chem. Commun. 2015, 51, 9010. (10) (a) Fukuzumi, S.; Kishi, T.; Kotani, H.; Lee, Y.-M.; Nam, W. Highly Efficient Photocatalytic Oxygenation Reactions Using Water as an Oxygen Source. Nat. Chem. 2011, 3, 38. (b) Fukuzumi, S.; Mizuno, T.; Ojiri, T. Catalytic Electron-Transfer Oxygenation of Substrates with Water as an Oxygen Source Using Manganese Porphyrins. Chem. - Eur. J. 2012, 18, 15794. (11) (a) Chen, G.; Chen, L.; Ma, L.; Kwong, H.-K.; Lau, T.-C. Photocatalytic Oxidation of Alkenes and Alcohols in Water by a Manganese(V) Nitrido Complex. Chem. Commun. 2016, 52, 9271. (b) Kwong, H.-K.; Lo, P.-K.; Lau, K.-C.; Lau, T.-C. Epoxidation of Alkenes and Oxidation of Alcohols with Hydrogen Peroxide Catalyzed by a Manganese(V) Nitrido Complex. Chem. Commun. 2011, 47, 4273. (12) Li, F.; Yu, M.; Jiang, Y.; Huang, F.; Li, Y.; Zhang, B.; Sun, L. Chemical and Photochemical Oxidation of Organic Substrates by Ruthenium Aqua Complexes with Water as an Oxygen Source. Chem. Commun. 2011, 47, 8949. (13) Iali, W.; Lanoe, P.-H.; Torelli, S.; Jouvenot, D.; Loiseau, F.; Lebrun, C.; Hamelin, O.; Ménage, S. A Ruthenium(II)-Copper(II) Dyad for the Photocatalytic Oxygenation of Organic Substrates Mediated by Dioxygen Activation. Angew. Chem., Int. Ed. 2015, 54, 8415. (14) (a) Sankaralingam, M.; Palaniandavar, M. Tuning the Olefin Epoxidation by Manganese(III) Complexes of Bisphenolate Ligands: Effect of Lewis Basicity of Ligands on Reactivity. Dalton Trans. 2014, 43, 538. (b) Saravanan, N.; Sankaralingam, M.; Palaniandavar, M. 5103

DOI: 10.1021/acs.inorgchem.7b00220 Inorg. Chem. 2017, 56, 5096−5104

Article

Inorganic Chemistry (25) (a) Duelund, L.; Hazell, R.; McKenzie, C. J.; Nielsen, L. P.; Toftlund, H. Solid and Solution State Structures of Mono-and DiNuclear Iron(III) Complexes of Related Hexadentate and Pentadentate Aminopyridyl Ligands. J. Chem. Soc., Dalton Trans. 2001, 152. (b) Sankaralingam, M.; Balamurugan, M.; Palaniandavar, M.; Vadivelu, P.; Suresh, C. H. Nickel(II) Complexes of Pentadentate N5 Ligands as Catalysts for Alkane Hydroxylation by Using m-CPBA as Oxidant: A Combined Experimental and Computational Study. Chem. - Eur. J. 2014, 20, 11346. (26) Nam, W.; Lee, Y.-M.; Fukuzumi, S. Tuning Reactivity and Mechanism in Oxidation Reactions by Mononuclear Nonheme Iron(IV)-Oxo Complexes. Acc. Chem. Res. 2014, 47, 1146. (27) de Oliveira, F. T.; Chanda, A.; Banerjee, D.; Shan, X.; Mondal, S.; Que, L., Jr.; Bominaar, E. L.; Münck, E.; Collins, T. J. Chemical and Spectroscopic Evidence for an FeV-Oxo Complex. Science 2007, 315, 835. (28) Ghosh, M.; Singh, K. K.; Panda, C.; Weitz, A.; Hendrich, M. P.; Collins, T. J.; Dhar, B. B.; Gupta, S. S. Formation of a Room Temperature Stable FeV(O) Complex: Reactivity toward Unactivated C-H Bonds. J. Am. Chem. Soc. 2014, 136, 9524. (29) Kundu, S.; Thompson, J. V. K.; Ryabov, A. D.; Collins, T. J. On the Reactivity of Mononuclear Iron(V)-Oxo Complexes. J. Am. Chem. Soc. 2011, 133, 18546. (30) Bissette, A. J.; Fletcher, S. P. Mechanisms of Autocatalysis. Angew. Chem., Int. Ed. 2013, 52, 12800. (31) Bissette, A. J.; Odell, B.; Fletcher, S. P. Physical Autocatalysis Driven by a Bond-Forming Thiol-Ene Reaction. Nat. Commun. 2014, 5, 4607. (32) Bissette, A. J.; Fletcher, S. P. Novel Applications of Physical Autocatalysis. Origins Life Evol. Biospheres 2015, 45, 21. (33) Carnall, J. M. A.; Waudby, C. A.; Belenguer, A. M.; Stuart, M. C. A.; Peyralans, J. J.-P.; Otto, S. Mechanosensitive Self-Replication Driven by Self-Organization. Science 2010, 327, 1502. (34) Vidonne, A.; Philp, D. Making Molecules Make Themselves The Chemistry of Artificial Replicators. Eur. J. Org. Chem. 2009, 2009, 593. (35) Semenov, S. N.; Kraft, L. J.; Ainla, A.; Zhao, M.; Baghbanzadeh, M.; Campbell, V. E.; Kang, K.; Fox, J. M.; Whitesides, G. M. Autocatalytic, Bistable, Oscillatory Networks of Biologically Relevant Organic Reactions. Nature 2016, 537, 656. (36) Bachmann, P. A.; Luisi, P. L.; Lang, J. Autocatalytic SelfReplicating Micelles as Models for Prebiotic Structures. Nature 1992, 357, 57. (37) Cao, C.; Wang, Q.-B.; Tang, L.-J.; Ge, B.-Q.; Chen, Z.-X.; Deng, S.-P. Chain-Length-Dependant Autocatalytic Hydrolysis of Fatty Acid Anhydrides in Polyethylene Glycol. J. Phys. Chem. B 2014, 118, 3461. (38) Schaufelberger, F.; Ramström, O. Kinetic Self-Sorting of Dynamic Covalent Catalysts with Systemic Feedback Regulation. J. Am. Chem. Soc. 2016, 138, 7836. (39) Ohkubo, K.; Hirose, K.; Fukuzumi, S. Photooxygenation of Alkanes by Dioxygen with p-Benzoquinone Derivatives with High Quantum Yields. Photochem. Photobiol. Sci. 2016, 15, 731. (40) (a) Mair, R. D.; Graupner, A. J. Determination of Organic Peroxides by Iodine Liberation Procedures. Anal. Chem. 1964, 36, 194. (b) Fukuzumi, S.; Kuroda, S.; Tanaka, T. Flavin Analogue-Metal Ion Complexes Acting as Efficient Photocatalysts in the Oxidation of pMethylbenzyl Alcohol by Oxygen under Irradiation with Visible Light. J. Am. Chem. Soc. 1985, 107, 3020. (c) Fukuzumi, S.; Tahsini, L.; Lee, Y.-M.; Ohkubo, K.; Nam, W.; Karlin, K. D. Factors That Control Catalytic Two- versus Four-Electron Reduction of Dioxygen by Copper Complexes. J. Am. Chem. Soc. 2012, 134, 7025. (41) Singh, K. K.; Tiwari, M. K.; Ghosh, M.; Panda, C.; Weitz, A.; Hendrich, M. P.; Dhar, B. B.; Vanka, K.; Gupta, S. S. Tuning the Reactivity of FeV(O) toward C-H Bonds at Room Temperature: Effect of Water. Inorg. Chem. 2015, 54, 1535. (42) Hong, S.; Jang, S. J.; Cho, K.-B.; Nam, W. Intermetal Oxygen Atom Transfer from an FeVO Complex to a MnIII Complex: An Experimental and Theoretical Approach. Chem. Commun. 2016, 52, 12968.

(43) Stephens, H. N. Studies in Auto-Oxidation. V. The Induction Period in Auto-Oxidation. J. Am. Chem. Soc. 1936, 58, 219. (44) For radical chain reactions in autoxidation, see; Tomás, R. A. F.; Bordado, J. C. M.; Gomes, J. F. P. p-Xylene Oxidation to Terephthalic Acid: A Literature Review Oriented toward Process Optimization and Development. Chem. Rev. 2013, 113, 7421. (45) (a) Grela, M. A.; Coronel, M. E. J.; Colussi, A. J. Quantitative Spin-Trapping Studies of Weakly Illuminated Titanium Dioxide Sols. Implications for the Mechanism of Photocatalysis. J. Phys. Chem. 1996, 100, 16940. (b) Buettner, G. R. Spin Trapping: ESR Parameters of Spin Adducts. Free Radical Biol. Med. 1987, 3, 259. (46) Karoui, H.; Nsanzumuhire, C.; Le Moigne, F.; Hardy, M.; Siri, D.; Derat, E.; Rockenbauer, A.; Ouari, O.; Tordo, P. Synthesis and Spin-Trapping Properties of a Trifluoromethyl Analogue of DMPO: 5Methyl-5-trifluoromethyl-1-pyrroline N-Oxide (5-TFDMPO). Chem. Eur. J. 2014, 20, 4064. (47) Zalomaeva, O. V.; Trukhan, N. N.; Ivanchikova, I. D.; Panchenko, A. A.; Roduner, E.; Talsi, E. P.; Sorokin, A. B.; Rogov, V. A.; Kholdeeva, O. A. EPR Study on the Mechanism of H2O2-Based Oxidation of Alkylphenols over Titanium Single-Site Catalysts. J. Mol. Catal. A: Chem. 2007, 277, 185. (48) Honeywill, J. D.; Mile, B. The Use of the Tertiary Alkyl Tetraoxide-Peroxyl Equilibrium, ROOOOR ⇌ 2RO2•, as a Clean Source of Tertiary Alkyl Peroxyls. J. Chem. Soc., Perkin Trans. 2 2002, 569. (49) Fukuzumi, S.; Ono, Y. Determination of the Cross Propagation Rate Constants in the Autoxidation of Hydrocarbons by the Electron Spin Resonance Technique. J. Phys. Chem. 1977, 81, 1895. (50) Fukuzumi, S.; Ono, Y. Electron Spin Resonance and Kinetic Studies on the Liquid-Phase Autoxidation of Tetralin with Lead Dioxide. J. Phys. Chem. 1976, 80, 2973. (51) Luo, Y.-R. Handbook of Bond Dissociation Energies in Organic Compounds; CRC Press: New York, 2002. (52) Similar autocatalytic radical chain pathways have been reported for the overall O2 activation processes. See: (a) Nishida, Y.; Lee, Y.-M.; Nam, W.; Fukuzumi, S. Autocatalytic Formation of an Iron(IV)-Oxo Complex via Scandium Ion-Promoted Radical Chain Autoxidation of an Iron(II) Complex with Dioxygen and Tetraphenylborate. J. Am. Chem. Soc. 2014, 136, 8042. (b) Morimoto, Y.; Lee, Y.-M.; Nam, W.; Fukuzumi, S. An Autocatalytic Radical Chain Pathway in Formation of an Iron(IV)-Oxo Complex by Oxidation of an Iron(II) Complex with Dioxygen and Isopropanol. Chem. Commun. 2013, 49, 2500. (c) Comba, P.; Lee, Y.-M.; Nam, W.; Waleska, A. Catalytic Oxidation of Alkanes by Iron Bispidine Complexes and Dioxygen: Oxygen Activation versus Autoxidation. Chem. Commun. 2014, 50, 412. (53) Armarego, W. L. F.; Chai, C. L. L. Purification of Laboratory Chemicals, 6th ed.; Pergamon Press: Oxford, U.K., 2009. (54) Frimer, A. A. A Convenient Synthesis of Allylic Hydroperoxides. J. Org. Chem. 1977, 42, 3194.

5104

DOI: 10.1021/acs.inorgchem.7b00220 Inorg. Chem. 2017, 56, 5096−5104